Fact-checked by Grok 2 weeks ago

Seminorm

In mathematics, particularly in functional analysis, a seminorm on a real or complex vector space X is a function p: X \to [0, \infty) that satisfies two key properties: absolute homogeneity, p(\lambda x) = |\lambda| p(x) for all scalars \lambda and vectors x \in X, and the triangle inequality, p(x + y) \leq p(x) + p(y) for all x, y \in X. Unlike a norm, a seminorm does not require positive-definiteness, meaning p(x) = 0 may hold for some nonzero x, allowing for a nontrivial kernel consisting of all vectors mapped to zero. Seminorms play a fundamental role in the study of topological vector spaces (TVS), where a single seminorm induces a topology via neighborhoods of the form \{x \in X : p(x) < \epsilon\} for \epsilon > 0, generating a Hausdorff, translation-invariant structure if the seminorm separates points (i.e., acts as a norm). More generally, a family of seminorms \{p_i\}_{i \in I} on X defines a locally convex topology by taking as a subbasis the sets \{x \in X : p_i(x) < \epsilon\} for i \in I and \epsilon > 0, which is essential for analyzing convergence, continuity, and completeness in infinite-dimensional spaces without a single dominating norm. This framework underpins the theory of locally convex spaces, including Fréchet and LF-spaces, and connects to broader concepts like the Hahn-Banach theorem via associated sublinear functionals. A notable application of seminorms arises in quotient constructions: for a seminorm p on X, the kernel N = \{x \in X : p(x) = 0\} forms a subspace, and the quotient space X/N inherits a genuine norm \|\overline{x}\| = p(x) (where \overline{x} = x + N), transforming the seminormed space into a normed one. Examples include the seminorm p(f) = \sup |f'(t)| on the space of continuously differentiable functions on [0,1], where constant functions lie in the kernel, or Minkowski functionals of absorbing, balanced convex sets, which yield seminorms under suitable conditions. These structures are crucial for applications in distribution theory, operator algebras, and partial differential equations, where seminorms facilitate the handling of topologies coarser than those induced by norms.

Fundamentals

Definition

A seminorm on a V over the real or complex numbers is a p: V \to [0, \infty) that satisfies the following three axioms for all x, y \in V and scalars \lambda \in \mathbb{R} or \mathbb{C}:
  • Non-negativity: p(x) \geq 0,
  • Absolute homogeneity: p(\lambda x) = |\lambda| p(x),
  • Subadditivity: p(x + y) \leq p(x) + p(y).
Unlike a norm, a seminorm does not require positive definiteness, meaning that p(x) = 0 need not imply x = 0; consequently, the kernel \{x \in V \mid p(x) = 0\} is a subspace of V that may be nontrivial. From the axioms, several properties follow. First, p(0) = 0: taking \lambda = 0 in the homogeneity axiom yields p(0 \cdot x) = |0| p(x) = 0, so p(0) = 0 for any x \in V. Second, p(-x) = p(x): by homogeneity with \lambda = -1, p(-x) = p((-1) x) = |-1| p(x) = p(x). The kernel is indeed a subspace: if p(x) = p(y) = 0, then p(x + y) \leq p(x) + p(y) = 0, so p(x + y) = 0 by non-negativity; similarly, p(\lambda x) = |\lambda| p(x) = 0.

Examples

A fundamental example of a seminorm is the trivial seminorm on any vector space X, defined by p(x) = 0 for all x \in X. This function satisfies subadditivity and absolute homogeneity, but its kernel is the entire space X, making it non-separating unless X = \{0\}. Every norm on a vector space is a seminorm, as norms satisfy the additional positive-definiteness property. For instance, the Euclidean norm on \mathbb{R}^n, given by \|x\|_2 = \sqrt{\sum_{i=1}^n x_i^2}, defines a seminorm whose kernel is only the zero vector. A specific illustration in \mathbb{R}^2 is the seminorm p(x, y) = |x|, which vanishes on the y-axis (the kernel consists of all vectors (0, y)) but is zero only for the origin along the x-axis direction. This example highlights how seminorms can identify subspaces where elements are indistinguishable. Seminorms arising from linear functionals provide another class: for a linear functional f: X \to \mathbb{R} (or \mathbb{C}), the function p(x) = |f(x)| is a seminorm, with kernel the annihilator of f. For example, on \mathbb{R}^n, if f(x) = \sum a_i x_i, then p(x) = |\sum a_i x_i| measures projection onto the direction defined by the coefficients a_i. In function spaces, seminorms often gauge behavior over domains. On the space C[a, b] of continuous functions on the compact interval [a, b], the supremum function p(f) = \sup_{t \in [a, b]} |f(t)| is a seminorm (in fact, a norm, as it separates points). The integral-based p(f) = \int_a^b |f(t)| \, dt similarly qualifies as a norm on this space. On the space C^1[a, b] of continuously differentiable functions, p(f) = \sup_{t \in [a, b]} |f'(t)| defines a seminorm whose kernel includes all constant functions. The sum of seminorms is itself a seminorm: if \{p_i\}_{i \in I} is a family of seminorms on X, then p(x) = \sum_{i \in I} p_i(x) (finite or absolutely convergent sums) satisfies the axioms, with kernel the intersection of the individual kernels. For example, on \mathbb{R}^2, p(x, y) = |x| + |y| combines componentwise absolute values.

Connections to Other Concepts

Minkowski Functionals

In , the Minkowski functional provides a geometric construction of seminorms from certain subsets of a . Consider a vector space V over the real or complex numbers equipped with a convex, balanced, and absorbing set B \subseteq V. The Minkowski functional associated to B is the function p_B: V \to [0, \infty] defined by p_B(x) = \inf \{ t > 0 \mid x \in tB \}, where tB = \{ t y \mid y \in B \} for t > 0, and the infimum over an empty set is taken to be \infty. This definition captures the "scaling factor" required to bring x into B, linking the intrinsic geometry of B to a functional perspective on V. The Minkowski functional p_B satisfies the axioms of a seminorm. Non-negativity holds because p_B(x) \geq 0 for all x \in V, as the infimum is over positive scalars, and finiteness follows from the absorbing property of B, which ensures that for every x \in V, there exists some t > 0 such that x \in tB. Absolute homogeneity, p_B(\alpha x) = |\alpha| p_B(x) for scalars \alpha, arises from the balanced condition on B, which implies |\alpha| B = B for |\alpha| = 1, allowing rescaling of the infimum. Subadditivity, p_B(x + y) \leq p_B(x) + p_B(y), follows from the convexity of B: if x \in t B and y \in s B, then x + y \in (t + s) B by the convex combination \frac{t}{t+s} \cdot \frac{x}{t} + \frac{s}{t+s} \cdot \frac{y}{s} \in B, so the infimum for x + y is at most t + s. Thus, under these conditions on B, p_B is a seminorm. Conversely, every seminorm on V arises as the Minkowski functional of a suitable set. For a seminorm p: V \to [0, \infty), the set K = \{ x \in V \mid p(x) \leq 1 \} is convex, as subadditivity and homogeneity imply that if p(x) \leq 1 and p(y) \leq 1, then p(\lambda x + (1 - \lambda) y) \leq \lambda p(x) + (1 - \lambda) p(y) \leq 1 for \lambda \in [0, 1]; balanced, by homogeneity with scalars of modulus at most 1; and absorbing, since for any x \in V with p(x) < \infty, x / (p(x) + 1) \in K if p(x) > 0, or x \in K if p(x) = 0. Moreover, p(x) = p_K(x) for all x \in V, establishing the bijection between seminorms and such sets (though K need not be bounded). If B is the closed unit ball of a \|\cdot\| on V, then the Minkowski functional p_B coincides exactly with \|\cdot\|, recovering the from its unit ball via the same infimum construction.

Relation to Norms and Pseudometrics

A p on a V over \mathbb{R} or \mathbb{C} is a if and only if it separates points, meaning p(x) = 0 implies x = 0. In this case, the condition ensures the structure aligns with that of a normed space, where the functional provides a complete when inducing a . The kernel of a seminorm p, defined as \ker(p) = \{ x \in V \mid p(x) = 0 \}, forms a of V. On the quotient space V / \ker(p), where elements are cosets = x + \ker(p), the functional p induces a via \| \| = p(x). This is well-defined because if = , then x - y \in \ker(p), so p(x) = p(y + (x - y)) \leq p(y) + p(x - y) = p(y), and symmetrically p(y) \leq p(x), implying equality; moreover, \| \| = 0 iff = {{grok:render&&&type=render_inline_citation&&&citation_id=0&&&citation_type=wikipedia}}, as p(x) = 0 means x \in \ker(p). Thus, every seminormed space quotients to a normed space under this construction. A seminorm p induces a pseudometric d on V by d(x, y) = p(x - y) for all x, y \in V. To verify d is a pseudometric, note non-negativity follows from p \geq 0; symmetry holds since d(y, x) = p(y - x) = p(-(x - y)) = | -1 | \, p(x - y) = d(x, y) by homogeneity; d(x, x) = p(0) = 0 since p(0) = p(2 \cdot 0) \leq 2 p(0) implies p(0) \leq 0, hence p(0) = 0; however, d(x, y) = 0 iff x - y \in \ker(p), which does not require x = y unless p is a norm; the triangle inequality follows from subadditivity: d(x, z) = p(x - z) = p((x - y) + (y - z)) \leq p(x - y) + p(y - z) = d(x, y) + d(y, z). Additionally, d is translation-invariant: d(x + a, y + a) = p((x + a) - (y + a)) = p(x - y) = d(x, y) for any a \in V. Pseudometrics generalize metrics in the same manner that seminorms generalize : both relax the separation of points, allowing distinct elements to have zero (or norm zero). In spaces derived from norms, zero distance implies equality, whereas pseudometric spaces permit non-trivial kernels, mirroring the structure in seminormed spaces. Bounded linear operators between seminormed spaces are defined analogously to those in normed spaces, with boundedness requiring \sup \{ p(Tx) / p(x) \mid x \in V \setminus \ker(p) \} < \infty, which transfers naturally to the quotient normed spaces.

Algebraic Properties

Basic Properties

Seminorms satisfy the property of sublinearity, meaning that for any vectors x, y in the vector space and any scalars \lambda, \mu, p(\lambda x + \mu y) \leq |\lambda| p(x) + |\mu| p(y). This follows directly from the axioms of positive homogeneity and subadditivity. To see this, first consider the real scalar case where \lambda, \mu \geq 0. By subadditivity, p(\lambda x + \mu y) \leq p(\lambda x) + p(\mu y), and by positive homogeneity, p(\lambda x) = \lambda p(x) and p(\mu y) = \mu p(y), yielding the inequality. For general real scalars, let \lambda' = |\lambda| and \mu' = |\mu|; then \lambda x = \operatorname{sign}(\lambda) (\lambda' x), and since p(\operatorname{sign}(\lambda) z) = p(z) for z = \lambda' x (as |\operatorname{sign}(\lambda)| = 1), the homogeneity extends to give p(\lambda x) = |\lambda| p(x). The same holds for \mu y, so subadditivity applies as above. For complex scalars, the absolute homogeneity p(\lambda x) = |\lambda| p(x) (where |\lambda| is the modulus) combines analogously with subadditivity to establish sublinearity, noting that multiplication by a complex number of modulus 1 preserves the seminorm value. The unit ball \{x \mid p(x) \leq 1\} of a p is convex. To prove this, take any x, y with p(x) \leq 1 and p(y) \leq 1, and any \lambda \in [0, 1]. Then, p(\lambda x + (1 - \lambda) y) \leq p(\lambda x) + p((1 - \lambda) y) = \lambda p(x) + (1 - \lambda) p(y) \leq \lambda \cdot 1 + (1 - \lambda) \cdot 1 = 1, where the first inequality is subadditivity and the equalities follow from positive homogeneity. Thus, \lambda x + (1 - \lambda) y lies in the unit ball, confirming convexity. Seminorms exhibit positive homogeneity, p(\lambda x) = \lambda p(x) for \lambda \geq 0, which implies a form of monotonicity. Specifically, for $0 \leq \lambda \leq 1, p(\lambda x) = \lambda p(x) \leq p(x), since \lambda \leq 1. This monotonicity reflects the scaling behavior inherent in the axioms. In the context of convex analysis, seminorms are precisely the sublinear functions that are non-negative and absolutely homogeneous, providing a bridge to the study of convex sets and functionals.

Inequalities

Seminorms satisfy the triangle inequality as a defining property: for a seminorm p on a vector space X, p(x + y) \leq p(x) + p(y) for all x, y \in X. This extends to finite sums by induction; for example, p(x_1 + \cdots + x_n) \leq p(x_1) + \cdots + p(x_n) for n \geq 1 and x_i \in X. The reverse triangle inequality holds: |p(x) - p(y)| \leq p(x - y) for all x, y \in X. To prove this, apply the triangle inequality to obtain p(x) = p((x - y) + y) \leq p(x - y) + p(y), so p(x) - p(y) \leq p(x - y). Similarly, p(y) = p((y - x) + x) \leq p(y - x) + p(x) = p(x - y) + p(x), yielding p(y) - p(x) \leq p(x - y). Combining these gives the result. A related bound is that for any \varepsilon > 0, \inf \{ p(y) \mid p(x - y) \leq \varepsilon \} \geq p(x) - \varepsilon, with the infimum also bounded above by p(x) + \varepsilon via the triangle inequality. This follows directly from the reverse triangle inequality applied to elements y satisfying p(x - y) \leq \varepsilon. Homogeneity for positive integers follows from the axioms. Specifically, p(nx) = n p(x) for n \in \mathbb{N} and x \in X. The proof proceeds by induction: the base case n=1 is trivial. Assuming it holds for n, then p((n+1)x) = p(nx + x) \leq p(nx) + p(x) = n p(x) + p(x) = (n+1) p(x). For the reverse, p(x) = p\left(\frac{1}{n} (nx)\right) = \frac{1}{n} p(nx) by absolute homogeneity, so p(nx) = n p(x). This extends to rational multiples: for q = m/n with m, n \in \mathbb{N}, p(q x) = p\left(\frac{m}{n} x\right) = \frac{1}{n} p(m x) = \frac{m}{n} p(x) = q p(x). In normed spaces, where seminorms are norms (positive definite), the inequalities reduce to the standard norm inequalities, including the triangle and reverse forms. For p-seminorms, such as those arising in L^p spaces with $1 \leq p \leq \infty, Hölder-like inequalities hold: if p and q are conjugate exponents ($1/p + 1/q = 1), then for suitable functions f, g, |\int f g| \leq \|f\|_p \|g\|_q, where \|\cdot\|_p denotes the L^p seminorm. The polar (or dual) seminorm p^* associated to a seminorm p on X is defined on the algebraic dual X' by p^*(f) = \sup \{ |f(x)| \mid x \in X, \, p(x) \leq 1 \}. This satisfies the seminorm axioms and measures the "size" of linear functionals relative to the unit ball of p.

Hahn-Banach Theorem

The Hahn-Banach extension theorem for seminorms states that if V is a vector space over the real or complex numbers, p: V \to [0, \infty) is a seminorm on V, M is a subspace of V, and f: M \to \mathbb{K} (where \mathbb{K} = \mathbb{R} or \mathbb{C}) is a linear functional satisfying |f(x)| \leq p(x) for all x \in M, then there exists a linear extension F: V \to \mathbb{K} such that F|_M = f and |F(x)| \leq p(x) for all x \in V. This holds because seminorms are sublinear functionals, satisfying p(x + y) \leq p(x) + p(y) and p(\lambda x) = |\lambda| p(x) for scalars \lambda, which aligns with the sublinearity required for the theorem. The proof proceeds by Zorn's lemma applied to the collection of all pairs (N, g), where N is a subspace containing M and g: N \to \mathbb{K} is linear with g|_M = f and |g(y)| \leq p(y) for y \in N, partially ordered by inclusion. A maximal element (V, F) yields the extension, as any proper subspace can be extended to a larger one by adjoining an element outside and choosing the value to preserve the inequality using the subadditivity of p. Alternatively, a constructive proof uses a Hamel basis of V relative to M, defining F coordinate-wise while ensuring the bound holds via sublinearity. For the complex case, reduce to the real case by considering the real part u = \operatorname{Re} f, extending it to a real-linear U with |U(z)| \leq p(z), then defining F(x) = U(x) - i U(i x), which satisfies |F(x)| \leq p(x). The core inequality is |F(x)| \leq p(x) \quad \forall x \in V, with equality on M. A key consequence is that the kernel of F contains the kernel of p, since if p(x) = 0, then |F(x)| \leq 0 implies F(x) = 0. The theorem applies directly to Minkowski functionals of convex, absorbing sets, which are sublinear (and seminorms if the set is balanced), enabling the construction of supporting hyperplanes: for a convex set C \subset V with nonempty interior and x_0 \notin C, the Minkowski functional p of C - y (for y \in \operatorname{int} C) allows extension of a suitable functional to separate x_0 - y from C - y by a hyperplane tangent to the boundary. Geometrically, this interprets as the separation of a point from a closed convex set using a hyperplane defined by the extended functional, where the seminorm bounds ensure the hyperplane supports the set without crossing it, providing a foundation for convex duality in spaces equipped with seminorms.

Topological Structure

Induced Pseudometric and Topology

Given a seminorm p on a vector space X over \mathbb{R} or \mathbb{C}, the function d_p: X \times X \to [0, \infty) defined by d_p(x, y) = p(x - y) for all x, y \in X satisfies the axioms of a pseudometric: it is symmetric, non-negative, and satisfies the triangle inequality d_p(x, z) \leq d_p(x, y) + d_p(y, z), but may have d_p(x, y) = 0 for distinct x, y. This equips X with the structure of a pseudometric space, where the pseudometric is translation-invariant due to the homogeneity and subadditivity of p. The pseudometric d_p induces a topology \tau_p on X, which is the coarsest topology making p continuous (equivalently, making the pseudometric d_p continuous). A local basis for this topology at any point x \in X consists of the open pseudometric balls V(x, \epsilon) = \{ y \in X \mid p(y - x) < \epsilon \}, \quad \epsilon > 0. These neighborhoods are translation-invariant, and the sets U_\epsilon = \{ y \in X \mid p(y) < \epsilon \} form a local basis at the origin, with the full basis given by their translates x + U_\epsilon. The topology \tau_p is Hausdorff if and only if p separates points (i.e., p(x) = 0 implies x = 0), in which case p is a norm and d_p is a metric. The open sets U_\epsilon are convex, balanced (absolutely convex), and absorbing: for any x \in X, there exists \lambda > 0 such that \lambda x \in U_\epsilon, since p(\mu x) = |\mu| p(x) allows scaling \mu < \epsilon / p(x) if p(x) > 0. Thus, \tau_p has a basis of convex, balanced, absorbing neighborhoods, making (X, \tau_p) a locally convex topological vector space, where addition and scalar multiplication are continuous. The topology is uniformizable, arising from the uniform structure generated by the pseudometric entourages \{(x, y) \mid p(x - y) < \epsilon\}, and is compatible with the vector space operations in the sense that it preserves the algebraic structure while ensuring joint continuity of the bilinears defining the space. For a family \{p_i\}_{i \in I} of seminorms on X, the induced topology is the coarsest making each p_i continuous, with a basis consisting of finite intersections of sets \{ y \mid p_{i_k}(y - x) < \epsilon_k \}; this coincides with the product topology on the uniform structure but is strictly the initial topology on X with respect to the family.

Equivalent Seminorms

In the context of seminorms on a vector space X, one seminorm q is said to be stronger than another seminorm p if p(x) \leq q(x) for all x \in X. Conversely, p is weaker than q. Two seminorms p and q are equivalent, denoted p \sim q, if each is stronger than the other, meaning p(x) \leq q(x) and q(x) \leq p(x) for all x \in X. Equivalent seminorms induce the same topology on X, as their respective families generate identical open sets in the locally convex topology defined by the seminorm bases. A precise characterization of equivalence is that p \sim q if and only if there exist constants c, C > 0 such that c \, q(x) \leq p(x) \leq C \, q(x) for all x \in X. This boundedness ensures that the seminorms are topologically indistinguishable. To see this via open sets, consider the topologies \tau_p and \tau_q induced by p and q, respectively. These are the coarsest topologies making p and q continuous, with neighborhood bases at the origin given by \{x + \epsilon \{y : p(y) < 1\} : \epsilon > 0\} and similarly for q. The topologies coincide if and only if the open unit balls B_p = \{x : p(x) < 1\} and B_q = \{x : q(x) < 1\} satisfy that B_p contains some scalar multiple c B_q with c > 0, and vice versa B_q contains C B_p for some C > 0. Indeed, if c q(x) \leq p(x) \leq C q(x), then B_p \supseteq c B_q and B_q \supseteq (1/C) B_p, ensuring every basic open set in one topology contains a basic open set from the other, and vice versa. In normed spaces, where the seminorm is actually a norm, equivalent seminorms yield isomorphic topologies, preserving the vector space structure and convergence properties. In Fréchet spaces, which are complete metrizable locally convex spaces defined by a countable family of seminorms, any two countable families generating the same topology are equivalent in the sense that each seminorm in one family is bounded by a finite linear combination of seminorms from the other family. In 2025 developments within semi-Hilbert spaces—inner product spaces equipped with seminorms derived from semi-inner products—equivalent operator seminorms have been shown to preserve isometries, extending classical results on orthogonality preservation to broader operator classes.

Normability and Separability

In a seminormed space (X, p), where p is a seminorm on the vector space X, the space is normable if and only if p induces a norm, meaning the topology generated by p coincides with that of a normed space. This occurs precisely when p separates points, i.e., p(x) = 0 implies x = 0, or equivalently, the kernel of p is trivial: \ker(p) = \{0\}. In this case, the Minkowski functional associated with a suitable absorbing set yields a true . More generally, for a topological vector space (TVS), normability requires the topology to be induced by a single ; by Kolmogorov's criterion, a T_1 TVS is normable if and only if it admits a convex, topologically bounded neighborhood of the origin. A seminorm p on a vector space separates points if p(x) = 0 for all x \in X implies x = 0, ensuring the induced pseudometric distinguishes distinct elements. For a family of seminorms \{p_i\}_{i \in I} generating a topology on a TVS, the family jointly separates points if, for every nonzero x \in X, there exists some i such that p_i(x) > 0; this condition is equivalent to the topology being Hausdorff. A TVS is seminormable if its topology is generated by a (possibly ) family of seminorms. This holds the space is Hausdorff and locally , as locally topologies admit a basis of open sets whose Minkowski functionals are seminorms, and the Hausdorff property ensures separation of points. Barrelled spaces, defined as locally TVS where every closed balanced absorbing set (barrel) is a neighborhood of zero, are thus seminormable by virtue of their local . Similarly, Montel spaces—barrelled TVS in which closed bounded sets are compact—are locally and hence seminormable. Counterexamples arise in non-locally TVS, such as \ell^p(\mathbb{N}) for $0 < p < 1, where the topology cannot be generated by seminorms despite being a complete metrizable TVS. In Fréchet spaces, which are complete metrizable TVS generated by countable families of seminorms that separate points, recent studies examine linear isometries and their preservation of seminorm structures. Specifically, under conditions on the metric's growth (e.g., via absolutely continuous functions bounding the seminorm sequence), linear isometries preserve the individual seminorms defining the metric, thereby maintaining separability properties.

Continuity Properties

In a seminormed space (V, p), the seminorm p: V \to \mathbb{R} is continuous with respect to the topology induced by p, since p(x) = d(x, 0) where d(x, y) = p(x - y) is the associated pseudometric, making the identity map from (V, d) to \mathbb{R} continuous. This holds for any single seminorm, as the induced topology ensures that neighborhoods of zero translate directly to boundedness in p. For linear maps between seminormed spaces, a linear operator T: (V, p) \to (W, q) is continuous if and only if it is bounded, meaning \sup \{ q(Tx) \mid p(x) \leq 1 \} < \infty, or equivalently, there exists M < \infty such that q(Tx) \leq M p(x) for all x \in V. In the more general setting of locally convex spaces defined by families of seminorms \{p_i\}_{i \in I} on V and \{q_j\}_{j \in J} on W, the linear map T is continuous if and only if \sup_i p_i(x) \leq 1 implies \sup_j q_j(Tx) < \infty for all such balanced convex sets. In locally convex spaces, continuity of a linear map is equivalent to continuity at the origin, since the space is translation-invariant and addition is jointly continuous. Translations in such spaces are uniformly continuous, as the uniform structure induced by the seminorms ensures that the modulus of continuity for \tau_h(x) = x + h depends only on the difference x - y, independent of the base point. The Hahn-Banach theorem implies that continuous extensions of linear functionals on subspaces preserve boundedness with respect to a given seminorm, ensuring the extension remains continuous in the full space. Specifically, if a linear functional f on a subspace satisfies |f(y)| \leq p(y) for a seminorm p, the extension \tilde{f} satisfies |\tilde{f}(x)| \leq p(x) for all x, maintaining the bound and thus continuity in the induced topology.

Generalizations and Extensions

Quasi-Seminorms and p-Seminorms

A quasi-seminorm on a vector space X over the reals or complexes is a function p: X \to [0, \infty) that satisfies absolute homogeneity, p(\lambda x) = |\lambda| p(x) for all scalars \lambda and x \in X, and a relaxed triangle inequality, p(x + y) \leq K (p(x) + p(y)) for some constant K \geq 1 and all x, y \in X. This generalization allows for the study of topological vector spaces (TVS) that may not be locally convex, where standard seminorms fail to generate the topology sufficiently. Unlike seminorms, the constant K > 1 introduces a controlled deviation from subadditivity, enabling applications in spaces with weaker convexity properties. In contrast, a p-seminorm for $0 < p \leq 1 on a vector space X satisfies p(\lambda x) = |\lambda|^p p(x) for scalars \lambda and x \in X, along with the triangle inequality p(x + y) \leq p(x) + p(y) for all x, y \in X. When p = 1, this recovers the standard seminorm axioms. For p < 1, the structure can exhibit ultrametric properties, where p(x + y) \leq \max\{p(x), p(y)\} holds in certain cases, leading to non-Archimedean topologies useful in valuation theory and generalized metric spaces. When p=1, a prominent example of a seminorm arises in fractional Sobolev spaces W^{s,1}(\Omega), where the Gagliardo seminorm is defined as _{s,1} = \iint_{\Omega \times \Omega} \frac{|u(x) - u(y)|}{|x - y|^{n + s }} \, dx \, dy for $0 < s < 1, capturing the smoothness of functions via differences scaled by distance (with analogous definitions for 1 < p < ∞ using the p-power inside the integral and raising to 1/p). Quasi-seminorms play a key role in non-locally convex TVS, where they define topologies without relying on convex neighborhoods. Recent advancements leverage quasi-seminorms in regularization techniques for inverse problems, particularly \ell^p quasi-seminorms with $0 < p < 1 in total variation models, to promote sparsity while mitigating oversmoothing effects in sparse signal recovery from overdispersed data.

Families of Seminorms

A family of seminorms \{p_\alpha\}_{\alpha \in A} on a vector space X over \mathbb{R} or \mathbb{C} generates a locally convex topology on X, known as the locally convex topology induced by the family, which is the coarsest topology making each p_\alpha continuous. This topology has a basis of neighborhoods of the origin consisting of finite intersections of sets of the form \{x \in X \mid p_\alpha(x) < \varepsilon_\alpha\} for \varepsilon_\alpha > 0, equivalently expressed using finite suprema: sets where \sup_{\alpha \in F} p_\alpha(x) < \varepsilon for finite subsets F \subseteq A and \varepsilon > 0. The resulting space is Hausdorff if and only if the family separates points, meaning \bigcap_{\alpha \in A} \ker p_\alpha = \{0\}. When the family is countable and increasing (i.e., p_n \leq p_{n+1} for all n), the induced topology is metrizable, with a translation-invariant metric given by d(x, y) = \sum_{n=1}^\infty 2^{-n} \frac{p_n(x - y)}{1 + p_n(x - y)}; if the space is complete under this metric, it is a Fréchet space. In contrast, uncountable families generate more general locally convex Hausdorff topologies on topological vector spaces (LC-TV spaces), where the neighborhood basis relies on finite suprema rather than a single metric. Two families of seminorms \{p_\alpha\} and \{q_\beta\} are equivalent if they induce the same on X, which occurs each p_\alpha is continuous with respect to the generated by \{q_\beta\} and vice versa; this equivalence preserves the locally structure and allows for the of a directed or saturated to simplify descriptions without altering the . Recent developments have applied families of uniformity seminorms on spaces of bounded sequences, such as \ell^\infty(\mathbb{Z}), to establish and theorems in the of multiple recurrence and convergence in , addressing previous gaps in non-commutative settings through recursive definitions of these seminorms on L^\infty(X).

Applications

In Functional Analysis and Operator Theory

In functional analysis, operator seminorms play a crucial role in studying bounded linear operators on Hilbert spaces, particularly through generalizations that connect numerical radii to specific seminorm structures. For an A-bounded operator Q on a complex Hilbert space (K, ⟨·,·⟩), where A is a positive operator, the seminorm Q_{A,m,\lambda,f} is defined to link the A-numerical radius with the operator A-seminorm, providing bounds and inequalities that refine classical estimates for operator behavior. This construction allows for the analysis of semi-Hilbertian operators, where the seminorm facilitates inequalities involving products and sums, enhancing understanding of operator norms in non-standard inner product settings. The Hahn-Banach theorem extends to dual spaces equipped with seminorms, enabling the prolongation of linear functionals that are bounded by a given seminorm from a subspace to the entire space while preserving the bound. In locally convex spaces defined by a family of seminorms, this extension ensures the functional remains continuous with respect to the topology induced by the seminorms, which is essential for constructing dual spaces and separation theorems in operator theory. Advancements in 2025 have introduced n-tuple operator seminorms that generalize the A-Euclidean operator radius for n-tuple bounded linear operators on complex Hilbert spaces, offering improved estimates for the A-joint numerical radius and applications to multi-operator inequalities. These seminorms preserve key properties related to isometries, extending classical radius concepts to higher-dimensional operator tuples and aiding in the study of joint spectra and stability. In partial differential equations (PDEs), Sobolev seminorms quantify the smoothness of functions via weak derivatives, with the k-th order seminorm defined as |u|_{H^k} = \left( \sum_{|\alpha|=k} \| D^\alpha u \|_{L^2}^2 \right)^{1/2}, where D^\alpha u are the weak partial derivatives of multi-index \alpha. This seminorm is pivotal in Sobolev embedding theorems, which map functions from Sobolev spaces into continuous or Hölder spaces, providing compactness and regularity results crucial for solving elliptic and parabolic PDEs. Developments since 2024 have explored seminorms in the context of Fréchet isometries on spaces metrized by sequences of seminorms, examining how distances derived from these seminorms influence linear isometries and their preservation under topological mappings. This work relates the metric structure to seminorm families, yielding applications to isometric embeddings and the classification of Fréchet spaces in operator theory.

In Ergodic Theory and Optimization

In , seminorms play a crucial role in analyzing multiple ergodic averages along , particularly in identifying factors that determine . For systems with transformations, explicit factors have been constructed using these seminorms to establish criteria for the of such averages. Specifically, the Gowers-Host-Kra seminorms provide bounds on the uniformity of functions, enabling proofs of for sequences of iterates in actions of \mathbb{Z}^d. These seminorms facilitate the of ergodic systems into factors where the averages behave predictably, as demonstrated in studies of multiple averages on amenable groups. Uniformity seminorms on \ell^\infty(\mathbb{Z}) extend this framework by quantifying the of bounded sequences with nilsequences, supporting and inverse theorems in additive combinatorics and ergodic theory. These seminorms, defined through multi-linear averages, yield a structure theorem that decomposes functions with large uniformity norm into nilsequences plus structured error terms. The norms bound correlations precisely, allowing inverse results that characterize high uniformity as arising from phases on nilmanifolds. Such tools have applications in proving multiple recurrence for actions, bridging combinatorial and dynamical perspectives. A 2018 survey on modern regularization methods for inverse problems discusses the use of seminorms within regularization schemes to mitigate oversmoothing, preserving edges and high-frequency details in reconstructions. It highlights how operator-dependent seminorms in Tikhonov frameworks balance data fidelity with sparsity-promoting penalties, achieving stable convergence for noisy data without excessive smoothing. These methods, often incorporating fractional Sobolev seminorms, ensure robustness in applications like imaging and signal processing.